23.12.2014 Views

Asymptotic Analyses for Atmospheric Flows and ... - FU Berlin, FB MI

Asymptotic Analyses for Atmospheric Flows and ... - FU Berlin, FB MI

Asymptotic Analyses for Atmospheric Flows and ... - FU Berlin, FB MI

SHOW MORE
SHOW LESS

You also want an ePaper? Increase the reach of your titles

YUMPU automatically turns print PDFs into web optimized ePapers that Google loves.

Klein, R.: Atmosphere <strong>Asymptotic</strong>s <strong>and</strong> Numerics – AUTHOR’S UPDATED PRINT 765<br />

ZAMM · Z. angew. Math. Mech. 80 (2000) 11-12,765–785<br />

Klein, R.<br />

<strong>Asymptotic</strong> <strong>Analyses</strong> <strong>for</strong> <strong>Atmospheric</strong> <strong>Flows</strong> <strong>and</strong> the Construction<br />

of <strong>Asymptotic</strong>ally Adaptive Numerical Methods<br />

Pr<strong>and</strong>tl’s boundary layer theory may be considered one of the origins of systematic scale analysis <strong>and</strong> asymptotics<br />

in fluid mechanics. Due to the vast scale differences in atmospheric flows such analyses have a particularly strong<br />

tradition in theoretical meteorology. Simplified asymptotic limit equations,derived through scale analysis,yield<br />

a deep insight into the dynamics of the atmosphere. Due to limited capacities of even the fastest computers,<br />

the use of such simplified equations has traditionally been a necessary precondition <strong>for</strong> successful approaches to<br />

numerical weather <strong>for</strong>ecasting <strong>and</strong> climate modelling.<br />

In the face of the continuing increase of available compute power there is now a strong tendency to relax as<br />

many simplifying scaling assumptions as possible <strong>and</strong> to go back to more complete <strong>and</strong> more complex balance<br />

equations in atmosphere flow computations. However,the simplified equations obtained through scaling analyses<br />

are generally associated with singular asymptotic limits of the full governing equations,<strong>and</strong> this has important<br />

consequences <strong>for</strong> the numerical integration of the latter. In these singular limit regimes dominant balances of a<br />

few terms in the governing equations lead to degeneracies <strong>and</strong> singular changes of the mathematical structure<br />

of the equations. Numerical models based on comprehensive equation systems must simultaneously represent<br />

these dominant balances <strong>and</strong> the subtle,but important,deviations from them. These requirements are partly in<br />

contradiction,<strong>and</strong> this can lead to severe restrictions of the accuracy <strong>and</strong>/or efficiency of numerical models.<br />

The present paper makes a case <strong>for</strong> a somewhat unconventional use of the results of scale analyses <strong>and</strong> multiple<br />

scales asymptotics. It demonstrates how,through the judicious implementation of asymptotic results,numerical<br />

discretizations of the full governing equations can be designed so that they operate with uni<strong>for</strong>m accuracy <strong>and</strong><br />

efficiency even when a singular limit regime is approached.<br />

Keywords: Multiple Scales <strong>Asymptotic</strong>s,<strong>Atmospheric</strong> <strong>Flows</strong>,Numerical Methods,<strong>Asymptotic</strong> Adaptivity<br />

1 Introduction<br />

Pr<strong>and</strong>tl’s boundary layer theory may be considered one of the origins of systematic scale analysis <strong>and</strong> asymptotics<br />

in fluid mechanics, (see, e.g., [34, 41, 36]). Due to the vast scale differences in atmospheric flows, such analyses<br />

have a particularly strong tradition in theoretical meteorology. Classical textbooks <strong>and</strong> research monographs (see,<br />

e.g., [17, 32, 44]) explain how scaling arguments, in combination with order-of-magnitude-estimates of observed<br />

data, may be used to deduce simplified equation systems that capture the essence of atmospheric flows, while


766 ZAMM · Z. angew. Math. Mech. 80 (2000) 11-12<br />

suppressing unnecessary details. These systems often allow a more comprehensive underst<strong>and</strong>ing of many observed<br />

phenomena, because they contain only those physical interactions that are truly relevant.<br />

These limit equations generally emerge because of a singularity of the governing equations that is associated<br />

with the particular flow regime considered. This aspect is important in the context of numerical modelling based<br />

on the full governing equations. <strong>Flows</strong> that are generally within or close to the regime of validity of the asymptotic<br />

limit equations will be governed essentially by the dominant balances revealed by the asymptotics. Yet, a flow<br />

solver designed <strong>for</strong> the full governing equations will generally not be able to properly represent these <strong>and</strong>, at the<br />

same time, capture the remaining higher order effects. A prominent example, which is also relevant to atmospheric<br />

flow modelling, is the numerical simulation of low Mach number flows using fully compressible flow solvers. It has<br />

been shown, e.g., in [42, 38, 18], that st<strong>and</strong>ard compressible flow solvers become inaccurate or even fail completely<br />

<strong>for</strong> Mach numbers smaller than about M = 10 −2 .<br />

In the remainder of this paper we will recount in section 2aversion of the full three-dimensional compressible<br />

flow equations <strong>for</strong> flow in an atmospheric layer on the rotating earth in non-dimensional <strong>for</strong>m. In section 3 we discuss<br />

qualitatively three singular asymptotic limit regimes that are relevant <strong>for</strong> different atmospheric flow applications,<br />

namely the quasi-geostrophic regime <strong>for</strong> synoptic scale flows (≈ 1000km), <strong>and</strong> the pseudo-incompressible <strong>and</strong><br />

anelastic regimes <strong>for</strong> mesoscale flows on scales less than or equal to the pressure scale height (i.e., ≤ 10km).<br />

Section 4then provides some of results from (multiple scales) asymptotics of atmospheric flows.<br />

The scalings<br />

<strong>for</strong> this analysis are chosen so as to match the relevant numerically resolved scales in modern numerical weather<br />

<strong>for</strong>ecast systems or regional climate models. Various issues regarding the numerical modelling of such flows on the<br />

basis of the fully three-dimensional compressible flow equations will be explained in section 5. A strategy <strong>for</strong> the<br />

design of suitable numerical schemes, which has been adopted in the author’s earlier work on low Mach number<br />

flows, will be summarized briefly. Section 6 provides conclusions <strong>and</strong> an outlook to future activities.<br />

2 Non-dimensional compressible flow equations in a rotating frame<br />

The discussions in this paper will be based on the compressible non-homentropic three dimensional flow equations<br />

in a rotating frame of reference (see, e.g., [17, 32, 44]):<br />

Sr ρ t + ∇·(ρ⃗v) = 0,<br />

Sr ⃗v t + ⃗v ·∇⃗v + 1 Ω<br />

Ro ⃗ × ⃗v + 1 1<br />

M 2 ρ ∇p + 1 ⃗ Fr 2 k = D ⃗ v ,<br />

Sr p t + ⃗v ·∇p + γp∇·⃗v = D p .<br />

(1)<br />

Let ρ ref ,v ref ,p ref denote characteristic values of density, flow velocity, <strong>and</strong> pressure, let l ref ,t ref be characteristic<br />

length <strong>and</strong> time scales <strong>for</strong> the considered flow, <strong>and</strong> let Ω = 1<br />

day<br />

be the earth rotation frequency, <strong>and</strong> g the modulus


Klein, R.: Atmosphere <strong>Asymptotic</strong>s <strong>and</strong> Numerics – AUTHOR’S UPDATED PRINT 767<br />

of the gravitational acceleration. Then the characteristic non-dimensional numbers Fr, M, Ro, <strong>and</strong> Sr are defined<br />

<strong>and</strong> labelled as<br />

Fr =<br />

v ref<br />

√<br />

glref<br />

Froude Number<br />

M =<br />

v ref<br />

√<br />

pref /ρ ref<br />

Mach Number<br />

Ro =<br />

Sr =<br />

v ref<br />

2Ωl ref<br />

l ref /t ref<br />

v ref<br />

Rossby Number<br />

Strouhal Number<br />

Furthermore, in (1), ρ, ⃗v, p are the non-dimensional density, velocity <strong>and</strong> pressure, t denotes time, <strong>and</strong> Ω, ⃗ ⃗ k are<br />

unit vectors pointing in the directions of the earth rotation axis <strong>and</strong> of the gravitational acceleration, respectively.<br />

⃗D v <strong>and</strong> D p are associated with the effects of viscous (turbulent) friction, dissipation, heat conduction, <strong>and</strong> radiation<br />

on the balances of momentum <strong>and</strong> energy. D p may also include effects of latent heat release, even though in this<br />

case the equation system would have to be extended to include the transport of the water phases. γ = c p /c v is the<br />

ratio of the specific heat capacities at constant pressure <strong>and</strong> at constant volume, respectively.<br />

In meteorological terminology, the homogeneous parts of (1) describe merely the “dynamics” of the atmosphere.<br />

In order to account <strong>for</strong> the so-called subgrid scale effects, suitable expressions representing turbulent<br />

transport, radiation effects, moisture transport, evaporation, condensation, precipitation, <strong>and</strong> possibly many other<br />

processes must be added to the “dynamic” part of the model. Such expressions are introduced in numerical atmospheric<br />

flow models, because it is generally not possible to resolve the atmospheric dynamics down to the very<br />

smallest flow scales by any realistic computational grid. There<strong>for</strong>e, the effects of “unresolved scales” on the resolved<br />

ones must be parameterized, <strong>and</strong> the according terms are termed “the physics” of the model. They are subsumed<br />

here in D ⃗ v <strong>and</strong> D p .<br />

In the following we concentrate on the dynamics only <strong>and</strong> we discuss various singular flow regimes that<br />

emerge when one or more of the non-dimensional characteristic numbers defined above become very small or very<br />

large.<br />

3 Examples of singular limit regimes in atmosphere flows<br />

3.1 Quasi-geostrophic flow<br />

One prominent example of a singular flow regime is the quasi-geostrophic limit, [17, 32, 44, 30], derived comprehensively<br />

by Pedlosky [32]. Following his exposition, we consider<br />

1. a shallow atmosphere, with a vertical extent h ref ≪ l ref ,


768 ZAMM · Z. angew. Math. Mech. 80 (2000) 11-12<br />

2. horizontal scales small compared with the earth radius, so the flow occurs approximately in a tangent plane,<br />

3. horizontal scales over which the vertical component of the earth rotation vector changes appreciably, so that<br />

⃗ k · ⃗ Ω=1+βy, where y is a cartesian coordinate pointing in the meridional direction, <strong>and</strong> β is a constant,<br />

4. small vertical velocities w so that w/v ref ≪ O(h ref /l ref ),<br />

5. low Mach, Froude, <strong>and</strong> Rossby numbers so that<br />

M → 0, Fr → 0, Ro → 0,<br />

with<br />

M M2<br />

=1, <strong>and</strong> → 0 as M → 0 , (2)<br />

Fr Ro<br />

6. <strong>and</strong> unsteady flows with Sr = O(1) as M, Fr, Ro → 0.<br />

At the leading order, the vertical momentum balance yields hydrostatics, <strong>and</strong> the horizontal momentum balance<br />

shows that the leading order pressure <strong>and</strong> density depend on the vertical coordinate z only. The Coriolis effect<br />

induces a perturbation of order<br />

δ = M2<br />

Ro ≪ 1 , (3)<br />

so that density <strong>and</strong> pressure may be exp<strong>and</strong>ed as<br />

p = p(z)+δ ρ(z) p ′ (⃗x, z, t)+o(δ),<br />

ρ = ρ(z)<br />

(<br />

)<br />

1+δρ ′ (⃗x, z, t)+o(δ) ,<br />

where<br />

dp<br />

dz = −ρ = −θ(z)p 1 γ . (4)<br />

Here ⃗x denotes the horizontal coordinates, <strong>and</strong> θ(z) is the leading order potential temperature distribution which<br />

must be considered an input to the problem <strong>for</strong>mulation. The velocity field is exp<strong>and</strong>ed as<br />

⃗v = ⃗u qg +Ro(⃗u (1) + w (1) ⃗ k)+o(Ro) , (5)<br />

where vectors ⃗u denote the horizontal velocity components <strong>and</strong> w ⃗ k the vertical one.<br />

<strong>Asymptotic</strong> expansions then yield a closed set of equations <strong>for</strong> the quasi-geostrophic horizontal velocity ⃗u qg ,<br />

the leading pressure perturbation p ′ <strong>and</strong> the potential temperature perturbation θ ′ = −ρ ′ + ρ<br />

γp (z) p′ :<br />

Horizontal momentum balance<br />

⃗ k × ⃗uqg + ∇ || p ′ =0 (6)<br />

Vertical momentum balance / hydrostatics<br />

θ ′ = ∂p′<br />

∂z<br />

Transport equation <strong>for</strong> the vertical vorticity ζ qg = ⃗ k · (∇×⃗u qg )<br />

( )(<br />

∂<br />

∂t + ⃗u qg ·∇ || ζ qg + βy + 1 ( ))<br />

∂ ρ<br />

ρ ∂z S θ′ = 1 (<br />

∂ ρD<br />

′ )<br />

p<br />

ρ ∂z γpS<br />

(7)<br />

(8)


Klein, R.: Atmosphere <strong>Asymptotic</strong>s <strong>and</strong> Numerics – AUTHOR’S UPDATED PRINT 769<br />

In these equations D ′ p is the suitably scaled source term from the original pressure evolution equation in (1), S is<br />

the static stability parameter defined by<br />

S = 1 θ<br />

dθ<br />

dz , (9)<br />

<strong>and</strong> ∇ ||<br />

abbreviates the horizontal components of the gradient operator.<br />

We notice that (6)–(9) is the first closed set of equations obtained in this asymptotic regime of low Mach, Froude,<br />

<strong>and</strong> Rossby numbers. While (6) <strong>and</strong> (7) are the leading order horizontal <strong>and</strong> vertical momentum equations the<br />

vorticity transport equation (8) is a result of higher order balances. Thus, flow simulations based on more comprehensive<br />

equation systems, such as the three-dimensional compressible flow equations, must use highly accurate<br />

numerical schemes that can reliably represent such higher order balances without explicitly extracting them through<br />

asymptotics.<br />

Furthermore, the fact that the velocity field is related to a scalar function p ′ as in (6) introduces a compatibility<br />

constraint <strong>for</strong> the horizontal velocity field. One cannot assign an arbitrary two-dimensional initial velocity field,<br />

but must ensure that ⃗u qg is divergence free to begin with. In addition, a meaningful computation must generally<br />

start with an almost balanced flow, so that unbalanced fast wave components do not pollute the solution.<br />

(Note, however, that Embid <strong>and</strong> Majda [13, 14] reveal a flow regime in which the long time average effects<br />

of fast gravity waves do not influence the underlying slow time scale balanced mean flow. Thus, the often cost<br />

intensive procedure of finding balanced initial data may sometimes be omitted.)<br />

3.2 The incompressible, pseudo-incompressible, <strong>and</strong> anelastic approximations<br />

High frequency acoustic waves are generally considered unimportant <strong>for</strong> the purposes of weather prediction or<br />

climate modelling. Here “high frequency” is meant to denote acoustic modes with wavelengths on the order of 10<br />

km or less <strong>and</strong> characteristic frequencies higher than 1 min −1 . A plausibility argument states that sizeable elastic<br />

perturbations cannot establish in the atmosphere, because acoustic waves rapidly redistribute the associated energy<br />

<strong>and</strong> lead to an equilibration void of any acoustic modes. This intuitive explanation may be quantified through an<br />

asymptotic analysis <strong>for</strong> vanishing Mach number,[36, 28], that has been backed up by rigorous justifications, e.g., in<br />

[11, 22, 37]. The resulting set of zero Mach number, variable density flow equations without gravity <strong>and</strong> rotation<br />

can be written as<br />

ρ t + ⃗v ·∇ρ = −ρ∇·⃗v,<br />

⃗v t + ⃗v ·∇⃗v + 1 ρ ∇p′ = ⃗ D v ,<br />

(10)<br />

∇·⃗v = D p<br />

γp ∞<br />

.


770 ZAMM · Z. angew. Math. Mech. 80 (2000) 11-12<br />

Here p ′ is a higher order perturbation pressure satisfying p = p+M 2 p ′ , with p ≡ const.. (Notice that <strong>for</strong> combustion<br />

in closed systems the background pressure p may be time dependent, but will always be spacially homogeneous.)<br />

The (elliptic) divergence constraint (10) 3 replaces the total energy or pressure evolution equation. It effectively<br />

suppresses acoustic modes <strong>and</strong> is there<strong>for</strong>e extremely important <strong>for</strong> numerical flow simulations: Fast acoustic<br />

wave propagation is eliminated <strong>and</strong> thus will neither lead to undesired time step restrictions, nor to unphysical<br />

oscillations that may occur when straight-<strong>for</strong>ward discretizations are used to represent hyperbolic equation systems<br />

(see, e.g., [27]). The constraint does induce a complication, in that it establishes an elliptic, Poisson-type equation<br />

<strong>for</strong> the pressure perturbation p ′ , namely<br />

which<br />

∇·( 1 ρ ∇p′ )=−∇ · (⃗v ·∇⃗v − ⃗ D v ) − ∂ ∂t<br />

( )<br />

Dp<br />

, (11)<br />

γp<br />

• is not trivially solved numerically, <strong>and</strong><br />

• has no connection with the thermodynamic equation of state which <strong>for</strong> compressible flows would establish a<br />

relation between the density <strong>and</strong> pressure fields.<br />

In practice, these complications are often considered less severe than the time step constraints associated with a<br />

proper representation of anyway unimportant acoustic waves.<br />

Durran [10] describes two similar velocity divergence constraints that are used frequently in the <strong>for</strong>mulation<br />

of numerical atmosphere flow models when the horizontal scales considered are comparable to the vertical ones<br />

(see also [17, 44, 12]). On such scales one may describe, e.g., the transport of pollutants in the atmosphere after<br />

they have exited a smoke stack (relevant length scales: ≈ 100 m – 1 km), or one may simulate the <strong>for</strong>mation <strong>and</strong><br />

evolution of convective clouds on scales comparable to the atmospheric scale height of about 10 km.<br />

The first constraint,<br />

∇·(ρ⃗v) =0 (12)<br />

leads to the “anelastic approximation”, [3, 31, 9]. The second yields the “pseudo-incompressible approximation”,<br />

[9], <strong>and</strong> reads<br />

∇·(ρθ⃗v) =<br />

D p<br />

γp γ−1<br />

γ<br />

. (13)<br />

Here<br />

θ = p1/γ<br />

ρ<br />

(14)<br />

is the st<strong>and</strong>ard “potential temperature” except <strong>for</strong> a constant factor. The potential temperature is a function of<br />

the thermodynamic entropy, (see, e.g., [17], section 3.7.3). The overbar denotes horizontal averaging at constant<br />

height, z = const..


Klein, R.: Atmosphere <strong>Asymptotic</strong>s <strong>and</strong> Numerics – AUTHOR’S UPDATED PRINT 771<br />

If either one of these constraints is adopted in a numerical model, the desired effect of suppression of acoustic<br />

modes is obtained. Yet, these constraints are not equivalent <strong>and</strong> which one best suits a given situation must be<br />

carefully evaluated. Estimates of the respective regimes of validity of the constraints in (12) <strong>and</strong> (13) as well as a<br />

discussion of consequences <strong>for</strong> numerical discretizations will be given in section 4.1 below. Here we merely re-iterate<br />

an interesting observation from [44, 5]:<br />

If the anelastic <strong>and</strong> the pseudo-incompressible approximations hold simultaneously, then either<br />

• the vertical flow velocity w satisfies the “diagnostic relation”<br />

or<br />

w = D ( )−1<br />

p 1 ∂θ<br />

, (15)<br />

γp θ ∂z<br />

• the potential temperature has an expansion θ = θ ∞ +M 2 θ (2) (⃗x, z, t), where θ ∞ ≡ const.<br />

Notice that the second option has been assumed by Ogura <strong>and</strong> Phillips [31] in their derivation of the anelastic<br />

approximation. (See also sections 45, 46 of [44].)<br />

4 <strong>Asymptotic</strong> analysis<br />

4.1 The anelastic <strong>and</strong> pseudo-incompressible divergence constraints<br />

Here we consider atmospheric motions on length scales of up to the pressure scale height, i.e.,<br />

l ref ≤ h sc ≈ 10km ,<br />

<strong>and</strong> we assume an isotropic scaling of characteristic lengths, i.e., no shallowness is assumed. Also we are interested<br />

in convective time scales, so that we choose<br />

t ref = l ref<br />

v ref<br />

<strong>and</strong> Sr ≡ 1 .<br />

The Rossby number may be estimated as<br />

Ro =<br />

v ref<br />

><br />

v ref<br />

≈ 5 ,<br />

2Ωl ref 2Ωh sc<br />

so that 1/Ro ≤ O(1) <strong>and</strong> the Coriolis effects are non-singular.<br />

To assess the relative order of magnitude of the Mach <strong>and</strong> Froude numbers we first observe that the Froude<br />

number based on the pressure scale height actually equals the Mach number. The pressure scale height may be<br />

assessed using the approximate hydrostatic balance ∂p/∂z ≈−ρg, <strong>and</strong> the scale height is defined as the vertical


772 ZAMM · Z. angew. Math. Mech. 80 (2000) 11-12<br />

distance over which the pressure changes appreciably, i.e., by an amount comparable to the reference pressure at<br />

the ground. There<strong>for</strong>e<br />

∣ ∣∣∣ ∂p<br />

h sc ∂z∣ ≈ h scρ ref g := p ref , or h sc g ≈ p ref<br />

.<br />

ρ ref<br />

Thus, from the definitions of the characteristic numbers we have<br />

Fr ≥ M (with Fr = M <strong>for</strong> l ref = h sc ) .<br />

The pseudo-incompressible approximation.<br />

It is generally observed that pressures in the troposphere<br />

at any given location do not change appreciably in comparison with the background pressure of about 1<br />

bar. Thus pressure variations are always small <strong>and</strong> we may write<br />

p = p(⃗x, z)+O(ε p ) where ε p ≪ 1 , (16)<br />

<strong>and</strong> p(⃗x, z) is a time independent local mean pressure. Also, high frequency pressure fluctuations on time scales<br />

much shorter than the convective time scales are either absent or of such low amplitudes that they do not appreciably<br />

affect the pressure time derivative. As a consequence we have<br />

∂p<br />

∂t = O(ε p,t) ≪ 1 . (17)<br />

With these estimates, the pressure evolution equation (1) 3 yields<br />

⃗v ·∇p + γp∇·⃗v = γp γ−1<br />

γ<br />

∇·(p 1 γ ⃗v) = Dp + O(ε p,t ) . (18)<br />

If the pressure perturbations of order O(ε p ) allow the additional estimate<br />

∇p = ∇p + O(ε p,x ) with ε p,x ≪ 1 , (19)<br />

we may neglect errors of order O(ε p ,ε p,t ,ε p,x ) to find the “generalized pseudo-incompressible constraint”<br />

∇·(p 1 γ ⃗v) =<br />

D p<br />

γp γ−1<br />

γ<br />

. (20)<br />

The only approximations that are needed to obtain this result concern the smallness of the overall pressure variations<br />

<strong>and</strong> of the pressure time <strong>and</strong> space derivatives.<br />

Thus this divergence constraint is a direct consequence of a<br />

degeneration of the pressure evolution, <strong>and</strong> not one of mass conservation (see also [23]). In contrast, the original<br />

pseudo-incompressible divergence constraint from (13) necessitates the additional requirement ρθ = p 1 γ . Unless<br />

one defines θ <strong>and</strong> ρ through<br />

θ = 〈θ〉 <strong>and</strong> ρ = 〈 1 ρ 〉−1 (21)<br />

with a suitable averaging operator〈·〉, this requirement will be satisfied only when fluctuations of potential temperature<br />

<strong>and</strong> density are small. This would not be the case, e.g., <strong>for</strong> hot exhaust gases exiting a smoke stack.


Klein, R.: Atmosphere <strong>Asymptotic</strong>s <strong>and</strong> Numerics – AUTHOR’S UPDATED PRINT 773<br />

The low Mach number limit <strong>for</strong> small scales. When l ref ≪ h sc , we have Fr ≫ M <strong>and</strong> an<br />

appropriate asymptotic pressure expansion reads<br />

p = p + ε Fr p (Fr) (⃗x, z, t)+M 2 p (2) (⃗x, z, t)+o(ε Fr , M 2 ) (22)<br />

where<br />

ε Fr = M2<br />

Fr 2 ≪ 1 . (23)<br />

The momentum equation at order O(M −2 ) reads<br />

∇p ≡ 0 , (24)<br />

so that p is a constant. For this case the generalized pseudo-incompressible constraint from (20) reduces to the<br />

st<strong>and</strong>ard divergence condition <strong>for</strong> non-adiabatic zero Mach number flow<br />

∇·⃗v = D p<br />

γp . (25)<br />

This constraint is common, e.g., in the theory of low Mach number combustion, [28].<br />

When Fr = O(1) we have ε Fr =M 2 <strong>and</strong> the pressure perturbations p (Fr) <strong>and</strong> p (2) need not be distinguished.<br />

In this regime, which describes flow fields on the 10 m scale, the leading order system of governing equations is that<br />

presented in (10), provided the gravitation term is included in the momentum source ⃗ D v . There are interesting<br />

intermediate regimes <strong>for</strong> which the condition in (23) holds, <strong>and</strong> <strong>for</strong> which there is a non-trivial distinguished limit<br />

Fr = M α with 0


774 ZAMM · Z. angew. Math. Mech. 80 (2000) 11-12<br />

The horizontal momentum balance at order O(M −2 ) yields<br />

∇ || p ≡ 0 or p = p(z) . (30)<br />

Here ∇ || abbreviates the horizontal components of the gradient operator. From (29) <strong>and</strong> the (assumed) time<br />

independence of p we conclude that<br />

ρ (0) ≡ ρ(z) . (31)<br />

With this result we obtain the anelastic divergence constraint (12) as the leading order continuity equation, i.e.,<br />

∇·(ρ⃗v (0) )=0. (32)<br />

We realize furthermore that all the conditions needed to justify the pseudo-incompressible divergence constraint<br />

from (20), namely p = p(⃗x, z)+ε p , ∂p/∂t = O(ε p,t ), <strong>and</strong> ∇p = ∇p + O(ε p,x ) are satisfied, with ε p = ε p,t =<br />

ε p,x =M 2 . Thus, the anelastic <strong>and</strong> the pseudo-incompressible divergence constraints hold simultaneously! Using<br />

(29), (30), <strong>and</strong> (31) we conclude, after some straight-<strong>for</strong>ward manipulations, that<br />

(<br />

)<br />

w (0) 1 dp 1 γ<br />

dz − 1 dρ<br />

= w (0) 1 dθ<br />

ρ dz θ dz = D(0) p<br />

γp , (33)<br />

p 1 γ<br />

where w (0) is the leading order vertical velocity <strong>and</strong> θ = p 1 γ /ρ.<br />

At this point we have to distinguish different<br />

regimes concerning the sign <strong>and</strong> order of magnitude of the dry adiabatic stability parameter<br />

1 dθ<br />

θ dz = S =: M2 N 2 , (34)<br />

where N is the — non-dimensional — “buoyancy frequency” or “Brunt-Väisälä” frequency. This quantity is a<br />

measure of the oscillation frequency of particles that are vertically displaced in a stably stratified atmosphere with<br />

dθ/dz > 0, (see, e.g., [17, 32, 44, 12]). For many, if not most, practical purposes one may assume stable stratification<br />

with N 2 > 0 in the troposphere <strong>and</strong> stratosphere. Pedlosky [32] provides an order of magnitude estimate (see his<br />

eq. (6.4.13)), which in the present notation amounts to<br />

M 2 N 2 = 1 θ<br />

dθ<br />

dz = O(10−1 ) . (35)<br />

With M ≈ 0.03, v ref = 10m/s, l ref = h sc =10 4 m,t ref = l ref /v ref this amounts to a dimensional frequency of<br />

N/t ref =10 −2 s −1 , (see also [32], Fig. 6.4.2, <strong>and</strong> [17], page 52).<br />

Stable Stratification: Assuming now that the stability parameter is non-zero <strong>and</strong> positive, the simultaneous<br />

realization of the anelastic <strong>and</strong> pseudo-incompressible divergence constraints from (33) leads us to a diagnostic,<br />

algebraic relation <strong>for</strong> the leading order vertical velocity,<br />

w (0) = D(0) p<br />

γp<br />

( 1<br />

θ<br />

)−1<br />

dθ<br />

dz<br />

if<br />

dθ<br />

dz > 0 , (36)


Klein, R.: Atmosphere <strong>Asymptotic</strong>s <strong>and</strong> Numerics – AUTHOR’S UPDATED PRINT 775<br />

as pointed out in section 1. Here is a summary of the relevant asymptotic limit equations <strong>for</strong> the horizontal velocity<br />

⃗u, the perturbation pressure p (2) , <strong>and</strong> the vertical velocity w in this regime.<br />

⃗u t + ⃗u ·∇ || ⃗u + w⃗u z + 1 ρ ∇ || p (2) = ⃗ D u ,<br />

∇ || · ⃗u = − 1 ∂ρw<br />

ρ ∂z ,<br />

( )−1<br />

w = D(0) p 1 dθ<br />

.<br />

γp θ dz<br />

(37)<br />

The energy source term D p may depend on w, ⃗u, ρ, p, <strong>and</strong> on additional variables, such as humidity, <strong>for</strong> which<br />

one would have to include additional inhomogeneous scalar transport equations. Since the purpose of the present<br />

derivations is to point out the structure of the low Mach / low Froude number singularities rather than being<br />

comprehensive, we omit detailled specifications of D p here.<br />

The equations in (37) are strongly degenerate, [44], in that the vertical momentum balance is no longer part<br />

of the prognostic set of evolution equations. In this regime, vertical motion can take place only if there is sufficiently<br />

strong heat input or heat loss through sources such as latent heat release of absorption of radiation. In the absence<br />

of heat exchange there is no vertical motion <strong>and</strong> the airflow occurs in vertically stacked, decoupled layers. One<br />

consequence of the algebraic constraint in (36), which is consistent with the discussions in [39], is that it generally<br />

<strong>for</strong>ces airflow around topographical obstacles rather than allowing flow over them.<br />

Nearly Neutral Stratification: The desire to model events of “deep convection” in well-mixed atmospheric<br />

boundary layers has led Ogura <strong>and</strong> Phillips [31] to introduce the assumption of near neutral stability with<br />

1 dθ<br />

θ dz = O(M2 ) . (38)<br />

With the dimensional parameters as in the previous paragraph, this corresponds to N/t ref ≈ 10 −3 s −1 , <strong>and</strong> thus<br />

to a much weaker stratification than be<strong>for</strong>e. Zeytounian ([44], chapter 45) provides a compact derivation of the<br />

relevant governing equations <strong>for</strong> the full three-dimensional velocity field, ⃗v, <strong>and</strong> the second order perturbation of<br />

the potential temperature, θ (2) . In our notation these equations read<br />

⃗v t + ⃗v ·∇⃗v + 1 ρ ∇p(2) = ⃗ D v + θ (2) ⃗g,<br />

θ (2)<br />

∇·(ρ⃗v) = 0,<br />

(39)<br />

t + ⃗v ·∇θ (2) = D(2) p<br />

γp .<br />

Importantly, in this regime the energy input represented by the source term D p must be small of order O(M 2 ), i.e.,<br />

D p =M 2 D (2)<br />

p + o(M 2 ) . (40)<br />

The leading order pressure <strong>and</strong> density stratifications now correspond to a neutrally stable atmosphere <strong>and</strong> are


776 ZAMM · Z. angew. Math. Mech. 80 (2000) 11-12<br />

explicitly given by<br />

p(z) =<br />

(<br />

1 − γ − 1<br />

γ<br />

) γ<br />

z<br />

γ−1<br />

,<br />

1<br />

ρ(z) =p(z)<br />

γ . (41)<br />

Zeytounian [44] discusses two-dimensional steady state solutions of the above equations. He reduces the equations<br />

to a single second order elliptic equation <strong>for</strong> a stream function ψ. This special case provides an excellent basis<br />

<strong>for</strong> the design of test problems <strong>for</strong> comprehensive solvers of the three dimensional anelastic weak stratification<br />

equations from (39).<br />

4.2 Long wave length dynamics on short convective time scales<br />

Thus far we have considered asymptotic flow regimes that were characterized by single length <strong>and</strong> time scales.<br />

We did account <strong>for</strong> shallow atmospheres with vertical extent small compared to the horizontal scales, but in each<br />

direction only a single characteristic scale has been accounted <strong>for</strong>. Furthermore, the reference time scale has been<br />

the characteristic time scale of convection, i.e., t ref<br />

= l ref /v ref . Meteorological applications, however, generally<br />

feature a multitude of scales.<br />

On the one h<strong>and</strong>, there are the “unresolved” or “subgrid scales”, which are smaller than what can be<br />

resolved by the computational grids of atmospheric flow models. The overall effects of these subgrid scales on the<br />

resolved ones is a very subtle <strong>and</strong> important issue. The complexity of the problem is even higher than that of<br />

classical turbulence modelling. Many of the simplifying assumptions that are often used in this latter area, such<br />

as homogeneous isotropic turbulence characteristics, are not valid in the atmosphere due to anisotropic scalings,<br />

the influence of gravity <strong>and</strong> stratification, etc.. In addition, there are microscale processes, e.g., in the context of<br />

cloud <strong>for</strong>mation, that can only be described by multi-phase fluid dynamics. We will not address the problems of<br />

“parametrization of subgrid scale effects” here.<br />

Another multiple scales aspect arises from the continuous increase of modern computational capacities. It<br />

is now possible to simultaneously resolve a multitude of length scales within a single computation. For example,<br />

the current generation of weather <strong>for</strong>ecast codes uses a horizontal grid spacing as small as ∼ 3km, while spanning<br />

overall horizontal distances of thous<strong>and</strong>s of kilometers. Some of the consequences of the simultaneous presence of<br />

multiple length scales can be explored using asymptotic multiple scales expansions, <strong>and</strong> this is the topic of the<br />

present section. The following discussions will be based on the detailed derivations in [5].<br />

During previous discussions we have seen already that the mathematical structure of the effective dynamical<br />

equations changes drastically with the considered length scales. When considering multiple length scales in a single<br />

solution, a subtle issue regarding the underlying characteristic time scales arises in addition: At low flow Mach<br />

numbers the characteristic time scale <strong>for</strong> convection is always much longer than that of acoustic wave propagation


Klein, R.: Atmosphere <strong>Asymptotic</strong>s <strong>and</strong> Numerics – AUTHOR’S UPDATED PRINT 777<br />

<strong>for</strong> a given length scale. If, however, the largest present scale, l max , exceeds the smallest one, l min , by a factor<br />

of 1/M or more, then the characteristic time of convection on the l min scale becomes comparable to the acoustic<br />

time scale <strong>for</strong> l max .<br />

It is then not at all clear that the “filtering of acoustic modes”, as implied, e.g., in the<br />

quasi-geostrophic, anelastic, <strong>and</strong> pseudo-incompressible approximations, should affect all flow scales, or whether<br />

it should be restricted to the smallest scales only. In fact, there are very long wavelength combined acoustic —<br />

gravitational wave modes, the so-called Lamb waves (see [17], p. 504–506), which are inherently compressible <strong>and</strong><br />

must be accounted <strong>for</strong> together with the quasi-incompressible small scale dynamics.<br />

Botta et al. [5] consider flows with a smallest scale comparable to the pressure scale height h sc <strong>and</strong> chose<br />

reference length <strong>and</strong> times <strong>for</strong> non-dimensionalization according to<br />

l ref = h sc ≈ 10 4 m ,<br />

t ref = l ref<br />

v ref<br />

≈ 10 3 s (42)<br />

(i.e., the reference flow velocity is v ref ≈ 10m/s). However, they are interested at the same time in much larger<br />

horizontal scales<br />

l ac = 1 M l ref ≈ 300km (43)<br />

<strong>for</strong> which one finds the acoustic time scale to match the convection time scale <strong>for</strong> l ref<br />

l ac<br />

= Ml ac<br />

= l ref<br />

= t ref . (44)<br />

c ref v ref v ref<br />

As discussed earlier, the Froude <strong>and</strong> Rossby numbers based on the reference scales may be assessed through the<br />

distinguished limits<br />

Ro = O(1) ,<br />

Fr<br />

= O(1) as M → 0 . (45)<br />

M<br />

Notice that the effective Rossby number <strong>for</strong> the larger “acoustic” scales becomes<br />

Ro ac = v ref<br />

Ωl ac<br />

= O(M) as M → 0 , (46)<br />

so that we may expect the effects of rotation to achieve increasing importance at the larger length scales.<br />

Following the low Mach number single time – multiple space scale analysis <strong>for</strong> weakly compressible flows from<br />

[23], Botta et al. then choose an asymptotic ansatz <strong>for</strong> solutions of (1) of the <strong>for</strong>m<br />

U(⃗x, z, t;M)=U (0) (⃗x, M⃗x, z, t)+MU (1) (⃗x, M⃗x, z, t)+M 2 U (2) (⃗x, M⃗x, z, t)+o(M 2 ) , (47)<br />

where U =(ρ, p, ⃗u, w) t denotes the solution vector. Without going into details we will summarize here some of the<br />

key results of the subsequent analysis. For future reference we note that horizontal spatial gradients <strong>for</strong> the above<br />

solution ansatz translate as follows<br />

∇ || U = ∑ ν<br />

M ν (∇ ⃗x +M∇ ⃗ξ ) U ν , (48)<br />

where ⃗ ξ =M⃗x.


778 ZAMM · Z. angew. Math. Mech. 80 (2000) 11-12<br />

Small scale dynamics:<br />

On the small scales comparable to h sc it is found that<br />

∇ ⃗x p (0) = ∇ ⃗ξ p (0) = ∇ ⃗x p (1) ≡ 0 so that p (0) = p(z,t) , p (1) = p (1) ( ⃗ ξ,z,t) . (49)<br />

Consistent with observations, temporal variations of the background pressure p are suppressed, so that p = p(z).<br />

Next one recovers the pseudo-incompressible flow equations <strong>for</strong> stratified flows from (37), except <strong>for</strong> one<br />

modification concerning the influence of long wavelength pressure gradients. The horizontal momentum equation<br />

now reads<br />

⃗u t + ⃗u ·∇ ⃗x ⃗u + w⃗u z + 1 ρ ∇ ⃗x p (2) = ⃗ D Ro<br />

u − 1 ρ ∇ ⃗ ξ<br />

p (1) , (50)<br />

where we have included the Coriolis terms in ⃗ D Ro<br />

u<br />

to abbreviate the notation. One may expect to find the fully<br />

three-dimensional anelastic approximation from (39), plus the long wave pressure gradient term as in (50), if<br />

variations of the potential temperature are assumed to be of order O(M 2 ) only. This case was not discussed in [5].<br />

The only new effect on the small scales is thus the pressure gradient term 1 ρ ∇ ξ ⃗ p (1) , which is responsible <strong>for</strong><br />

a bulk acceleration of the fluid. Notice that the ⃗x-dependence of the second order pressure p (2) is determined by a<br />

small scale divergence constraint analogous to (37) 2 .<br />

Large scale dynamics:<br />

In a st<strong>and</strong>ard fashion, the relevant equation <strong>for</strong> the dynamics on the 1 M l refscale<br />

are obtained by horizontally averaging over the small scale coordinate ⃗x. The ⃗x-averaging operator will be<br />

abbreviated by 〈·〉. The large scale dynamics involves three variables, namely the long wave pressure mode p (1) ,<br />

the averaged vertical component of vorticity ζ = ⃗ k ·∇ ⃗ξ<br />

×〈⃗u (0) 〉, <strong>and</strong> the averaged first order vertical velocity<br />

w (1) = 〈w (1) 〉. The long wavelength dynamical equations <strong>for</strong> the considered regime read<br />

( ∂2<br />

∂t 2 − c2 ∇ ⃗ξ 2 ) p (1) = (1− c 2 ∂ ∂z ) ∂ ∂t (ρ w(1) ) − c 2 ζ + Q p (1) ,<br />

N 2 c 2 (ρ w (1) ) = −(1 + c 2 ∂ ∂z ) ∂ ∂t<br />

p (1) + Q ρw (1) ,<br />

(51)<br />

∂<br />

∂t ζ = 1<br />

c 2 ∂<br />

∂t p(1) + ∂ ∂z (ρ w(1) ) + Q ζ<br />

.<br />

Here Q i are inhomogeneous terms that combine the effects of small scale averages of nonlinear terms, <strong>and</strong> of the<br />

momentum <strong>and</strong> energy inhomogeneities ⃗ D v ,D p from (1). The system in (51) supports internal gravity waves, long<br />

wave acoustics, <strong>and</strong> the Lamb wave. It represents the link between the multiple length scale asymptotics from [5]<br />

<strong>and</strong> classical analyses <strong>for</strong> small perturbations from a state of rest.<br />

We notice also that the in the stationary limit (∂/∂t ≡ 0) the first equation in (51) reduces to the pressure<br />

Poisson equation from the quasi-geostrophic theory obtained by applying the horizontal divergence ∇ ||<br />

· (·) to<br />

the geostrophic balance equation (6). Thus we verify that the Coriolis effects achieve dominant importance <strong>for</strong><br />

quasi-steady flow on the larger ⃗ ξ-scale as announced earlier.


Klein, R.: Atmosphere <strong>Asymptotic</strong>s <strong>and</strong> Numerics – AUTHOR’S UPDATED PRINT 779<br />

5 Numerical issues<br />

5.1 Discussion of the singular limit regimes<br />

In the previous sections we have provided a summary of some important singular limit regimes <strong>for</strong> atmospheric<br />

flows. Simplified asymptotic limit equations have been discussed that can be derived systematically from the fully<br />

three-dimensional compressible flow equations by considering various distinguished limits connecting the Mach,<br />

Froude, <strong>and</strong> Rossby numbers. (We have focused on convective time scales, so that the Strouhal number has been<br />

set to unity throughout). Based on these considerations we will now point out a number of difficulties that must be<br />

addressed when near-singular solutions to the original compressible flow equations are to be obtained via numerical<br />

integration.<br />

Low Mach number small scale flows:<br />

Consider first the regimes of pseudo-incompressible <strong>and</strong><br />

anelastic flows from section 3.2 <strong>and</strong> 4.1. It has been pointed out that the stratification of potential temperature<br />

plays an crucial role <strong>for</strong> the flow dynamics. Under near neutral conditions, where θ = θ ∞ +M 2 θ (2) , there is fully<br />

three-dimensional dynamics, whereas the vertical motion is strongly blocked by buoyancy <strong>for</strong>ces if the stratification<br />

is stronger. Furthermore, in this regime there is an asymptotic decoupling between the thermodynamic component<br />

of the pressure, the background pressure p, <strong>and</strong> the pressure gradients in the momentum equation, which are<br />

represented by the second order pressure p (2) .<br />

Suppose now that a fully compressible flow solver using pressure, density, <strong>and</strong> velocity as the dependent<br />

variables were used to compute a weakly stratified flow. Let us suppose further that a second order accurate discretization<br />

is used <strong>and</strong> that the numerical resolution is of the order of h = l ref /∆x ≈ 10 −2 , where ∆x characterizes<br />

the mesh size of the computational grid. Local truncation errors within the pressure <strong>and</strong> density evolution equations<br />

will then be of order δρ, δp = O(h 2 ) ∼ 10 −4 . After a limited number of time steps the potential temperature<br />

θ = p 1 γ /ρ will have accumulated errors of order O(10 −3 ), which is comparable to the square of the Mach number.<br />

Thus, unless additional special measures are taken, the buoyancy <strong>for</strong>ces in the vertical momentum equation will<br />

be dominated by the net effect of truncation errors, <strong>and</strong> an accurate prediction of vertical velocities cannot be<br />

expected.<br />

Suppose next that the same flow solver is used to compute a flow field with stable stratification as described<br />

by equations (37). The diagnostic constraint <strong>for</strong> the leading order vertical velocity states that, upon heat addition,<br />

a mass element will move vertically within the stratification so as to maintain a position of neutral buoyancy. This<br />

motion is a result of an intimate coupling that simultaneously involves the pressure, density, <strong>and</strong> vertical momentum<br />

equations. Unless specially designed to respect this balance, a fully compressible flow solver will respond to heat<br />

addition by generating unphysical acoustic waves with amplitudes of the order of the numerical discretization error.


780 ZAMM · Z. angew. Math. Mech. 80 (2000) 11-12<br />

Buoyancy imbalances induced by truncation errors can in addition excite unphysical internal gravity waves.<br />

Another deterioration of accuracy with decreasing Mach number is associated with the asymptotic pressure<br />

decomposition. Compressible flow solvers which feature a single pressure variable without explicitly implementing<br />

the asymptotic scalings have been observed to fail <strong>for</strong> sufficiently small Mach number in [6, 42, 38, 18]. This failure<br />

can be traced to back to cancellation of significant digits in the discretization of the pressure gradient. There is a<br />

large volume of literature addressing the issue of low Mach number flow numerics. Besides the references already<br />

cited here is an additional, un<strong>for</strong>tunately still incomplete, list <strong>for</strong> further reading, [1, 2, 4, 6, 16, 21, 23, 25, 26, 29, 40].<br />

(See also section 5.2 below.)<br />

In addition to these accuracy issues, compressible flow solvers often become inefficient at low Mach numbers,<br />

because they are subject to stability related time step restrictions associated with the acoustic wave speed. The<br />

regime of anelastic, weakly stratified flow, (39), allows the construction of more efficient schemes with time steps<br />

restricted by the magnitude of the convection velocity rather than the sound speed. The complexity of an associated<br />

computational model is comparable to that of a three-dimensional incompressible flow solver. If one is interested<br />

furthermore in flows with adiabatically stable stratification, (37), then the vertical velocity is given diagnostically<br />

<strong>and</strong> one is left with two-dimensional divergence constraints <strong>for</strong> the horizontal velocity in each horizontal layer of the<br />

computational grid. Imposing a set of independent horizontal divergence constraints should be much more efficient<br />

than a single three-dimensional one, <strong>and</strong> a considerable further gain in computational speed can potentially be<br />

achieved.<br />

Multiple scales:<br />

Large (synoptic) scale flows in the atmosphere may be approximated quite well<br />

by the quasi-geostrophic theory summarized in section 3.1. However, the quasi-geostrophic equations are highly<br />

filtered in that they suppress most of the wave phenomena in the atmosphere, such as long wave acoustic gravity<br />

waves <strong>and</strong> internal gravity waves. Furthermore, modern computational capacities allow the simultaneous numerical<br />

representation of such long wavelength phenomena <strong>and</strong> flow features on length scales of a few kilometres. We have<br />

seen that the anelastic <strong>and</strong> pseudo-incompressible flow limits, which are relevant <strong>for</strong> these smaller scales, are<br />

described by equation systems of a very different nature. It is a highly non-trivial challenge to construct numerical<br />

schemes which represent the almost balanced large scale flows, at the same time resolve the much richer small<br />

scale dynamics, <strong>and</strong> provide high accuracy on both. In the next section we briefly summarize a strategy <strong>for</strong> the<br />

construction of multiple scales numerics that combines ideas from asymptotic analysis with numerical multi-grid<br />

techniques.


Klein, R.: Atmosphere <strong>Asymptotic</strong>s <strong>and</strong> Numerics – AUTHOR’S UPDATED PRINT 781<br />

5.2 <strong>Asymptotic</strong>ally adaptive numerical methods<br />

Here we consider low Mach number flows without gravity <strong>and</strong> rotation that are characterized by a single time scale,<br />

but multiple space scale analogous to what has been presented in section 4.2. In [23] the author derives simplified<br />

asymptotic limit equations that describe interaction of the long wave acoustic <strong>and</strong> short wave quasi-incompressible<br />

flow phenomena. Using a solution ansatz analogous to that in (47) the author obtains the following leading order<br />

set of equations <strong>for</strong> inviscid, adiabatic flows:<br />

Small scale quasi-incompressible flow<br />

ρ (0)<br />

t + ∇ ⃗x · (ρ (0) ⃗v (0) ) = 0<br />

(ρ (0) ⃗v (0) ) t + ∇ ⃗x · (ρ (0) ⃗v (0) ◦ ⃗v (0) )+∇ ⃗x p (2) = −∇ ⃗ξ p (1)<br />

(52)<br />

∇ ⃗x · ⃗v (0) = 0<br />

Long wave acoustic modes<br />

ρ (0) t = 0<br />

(ρ (0) ⃗v (0) ) t<br />

+ ∇ ⃗ξ p (1) = 0<br />

(53)<br />

p (1)<br />

t + γP 0 ∇ ⃗ξ · ⃗v (0) = 0<br />

Here P 0 is the leading order pressure, which is constant in this regime. As in section (47), the first order pressure<br />

does not have small scale variation, so that ∇ ⃗x p (1) ≡ 0 <strong>and</strong> p (1) = p (1) ( ξ,t). ⃗<br />

Solution techniques <strong>for</strong> the full compressible flow equations from (1) are suggested in [15, 16, 23, 24, 33]<br />

which are designed to operate efficiently <strong>and</strong> accurately in the present asymptotic regime of weakly compressible<br />

low Mach number flows. The idea is to first obtain efficient <strong>and</strong> accurate estimates of the key physical effects in<br />

the flow by using appropriate, optimized numerical methods to integrate the asymptotic limit equations over one<br />

time step. Then, ideas similar to those behind projection methods <strong>for</strong> incompressible flows, [19, 7, 8], are used to<br />

compose a final solution that actually obeys the full governing equations instead of merely the asymptotic limit<br />

equations. Here is a qualitative description of the key steps need to address the flow regime described by (52),<br />

(53).<br />

In step 1 one obtains an explicit estimate of the effects of nonlinear convection using an upwind based higher<br />

order discretization. The time step restriction <strong>for</strong> this procedure is given by a stability constraint based on the<br />

maximum flow velocity (instead of on the sound speed).<br />

Step 2 consists of (i) a filtering technique that extracts the current long wave components of the pressure,<br />

momentum, <strong>and</strong> velocity fields, p (1) , ⃗v, ρ⃗v. These data are then used to solve integrate (53) over one time step


782 ZAMM · Z. angew. Math. Mech. 80 (2000) 11-12<br />

using a discretization that overcomes the sound speed based stability constraints. Here one may use an implicit<br />

discretization on the original mesh. Alternatively, one may consider the long wave – short wave decomposition as<br />

the basis of a physics-induced multi-grid technique <strong>and</strong> solve (53) by an explicit method on a coarse grid, [33].<br />

The final step determines the small scale, second order pressure p (2) , which is responsible <strong>for</strong> guaranteeing<br />

that the discrete flow field locally obeys a divergence constraint as in (52) 3 . This step corresponds to a “projection<br />

step” in classical incompressible flow solvers.<br />

Notice that steps 2 <strong>and</strong> 3 yield independent evaluations of discrete analogues of the first <strong>and</strong> second order<br />

pressures p (1) ,p (2) from the asymptotic analysis. In the numerical scheme, these pressure contributions are computed<br />

on the basis discretizations that are non-singular as the Mach number vanishes. Thus, the above-mentioned<br />

cancellation of significant digits is avoided. Only after completion of the time step are the pressure contributions<br />

re-composed to yield the full thermodynamic pressure.<br />

A subtle issue in this approach is the fact that the numerical discretization makes explicit use of the current<br />

value of the Mach number. It is thus necessary to implement intelligent filtering techniques that extract the long<br />

wave <strong>and</strong> short wave solution components <strong>and</strong>, simultaneously, simultaneously yield the instantaneous relevant<br />

value of the Mach number. This issue has been addressed in [24].<br />

Depending on the flow regimes that a computational code is mostly used <strong>for</strong>, the detailed components <strong>for</strong><br />

steps 1 to 3 may vary. The original idea in [23] was to extend a higher order shock capturing technique <strong>for</strong> fully<br />

compressible flows to the regime of small <strong>and</strong> even zero Mach number, <strong>and</strong> to include long wave acoustic effects.<br />

The desired extension to zero Mach number, variable density, multi-dimensional flows has been achieved in [35],<br />

code versions <strong>for</strong> flows including long wave acoustics have been presented in [15, 16].<br />

If one is interested mostly in flows at very low Mach number, then the overhead associated with full-fledged<br />

shock capturing techniques are unnecessary <strong>and</strong> simplifications can be introduced. Thus, Roller et al. [33] start from<br />

SIMPLE-type incompressible flow solvers, (see, e.g., [21]), <strong>and</strong> extend these technologies to the weakly compressible<br />

flow regime using the above techniques. Another alternative strategy has been proposed in by Worlikar et al. [43]<br />

in order to simulate the flow in a thermo-acoustic device. Their scheme is based on a vorticity stream function<br />

<strong>for</strong>mulation <strong>for</strong> the small scale quasi-incompressible flow, <strong>and</strong> uses the modified Godunov-Type scheme from [23]<br />

<strong>for</strong> long wave acoustics.<br />

6 Conclusions <strong>and</strong> outlook<br />

The “traditional” aim of scale analysis in meteorology is to obtain simplified asymptotic limit equations that are<br />

easier to solve than the more comprehensive fully compressible flow equations. There is a large number of such<br />

simplified model equations, <strong>and</strong> each has its particular advantages <strong>and</strong> shortcomings in practice. In recent years the


Klein, R.: Atmosphere <strong>Asymptotic</strong>s <strong>and</strong> Numerics – AUTHOR’S UPDATED PRINT 783<br />

rapid increase of compute power has stimulated strong ef<strong>for</strong>ts to relax the various simplifying scaling assumptions<br />

associated with asymptotic limit equations <strong>and</strong> to tackle the numerical solution of the full governing equations<br />

directly. It has become clear, however, that a lack of computing power has not been the only obstacle in attempts<br />

at solving these equations in earlier decades. Additional numerical issues arise, which are intimately related to<br />

the fact that scale analysis <strong>and</strong> asymptotic modelling have been successful in the first place. The derivation of<br />

simplified asymptotic models usually takes advantage of a singular degeneration of the governing equations in the<br />

respective limit regime, <strong>and</strong> it is these singularities that can induce severe numerical stiffnesses, <strong>and</strong> dynamic range<br />

problems.<br />

This paper advocates a so far relatively unconventional use of asymptotic scale analysis in the context of<br />

atmosphere flow modelling (see, however, [20]). We propose to construct new classes of “asymptotically adaptive<br />

numerical methods”, which do solve the full three dimensional compressible flow equations, but use the results<br />

of asymptotic scale analysis in the design of the discretizations.<br />

Such a scheme would assess a small number<br />

of nondimensional characteristic numbers “on the fly” during a computation. These characteristic numbers are<br />

chosen so as to indicate whether the current flow state is or is not within the vicinity of a singular limit regime.<br />

As a singular limit is approached, the discretizations automatically adapt <strong>and</strong> they merge into a scheme <strong>for</strong> the<br />

asymptotic limit equations when the limit is actually achieved.<br />

The strategy has been explained <strong>for</strong> the case of low Mach number multi-dimensional flows, <strong>for</strong> which it has<br />

been shown to be successful in a series of publications.<br />

Acknowledgement. The author gratefully acknowledges the continous constructive co-operation with Dr.<br />

Ann Almgren,Lawrence Berkeley National Laboratory,USA,Dr. Omar M. Knio,Johns Hopkins University,Baltimore,<br />

MD,USA,<strong>and</strong> Dr. Nicola Botta,Potsdam Institute <strong>for</strong> Climate Impact Research,Germany. This work has been sponsored<br />

partly by the Deutsche Forschungsgemeinschaft under Grant KL 611/6-1,2.<br />

References<br />

1 Abarbanel, S.; Duth, P.; Gottlieb, D.: Splitting method <strong>for</strong> low Mach number Euler <strong>and</strong> Navier-Stokes equations.<br />

Computers <strong>and</strong> Fluids, 17 (1989),1–12.<br />

2 Almgren, A.; Bell, J.; Colella, P.; Howell, L.; Welcome, M.: A conservative adaptive projection method <strong>for</strong><br />

the variable density incompressible navier-stokes equations. LBNL Preprint,(1996) 39075 UC-405.<br />

3 Batchelor, G. K.: The condition <strong>for</strong> dynamical similarity of motions of a frictionless perfect gas atmosphere. Quarterly<br />

Journal of Royal Meteorological Society, 79 (1953),224–235.<br />

4 Bijl, H.; Wesseling, P. A.: Unified method <strong>for</strong> computing incompressible <strong>and</strong> compressible flows in boundary-fitted<br />

coordinates. Journal of Computational Physics, 141 (1998),153173.


784 ZAMM · Z. angew. Math. Mech. 80 (2000) 11-12<br />

5 Botta, N.; Klein, R.; Almgren, A.: Dry atmosphere asymptotics. Techn. rep.,Potsdam Institute <strong>for</strong> Climate Impact<br />

Research,Potsdam,Germany,1999.<br />

6 Casulli, V.; Greenspan, D.: Pressure method <strong>for</strong> the numerical solution or transient compressible fluid flow. Int. J.<br />

Num. Meth. Fluids, 4 (1984),1001–1012.<br />

7 Chorin, A. J.: A numerical method <strong>for</strong> solving incompressible flow problems. J. Comput. Phys., 2 (1967),12–26.<br />

8 Chorin, A. J.: Numerical solution of the Navier-Stokes equations. Math. Comp., 22 (1968),745–762.<br />

9 Durran, D. R.: Improving the anelastic approximation. Journal of the <strong>Atmospheric</strong> Sciences, 46 (1989),1453–1461.<br />

10 Durran, D. R.: Numerical Methods <strong>for</strong> Wave Equations in Geophysical Fluid Dynamics. Springer,1999.<br />

11 Ebin, D. G.: Motion of Slightly Compressible Fluids in a Bounded Domain. i. Comm. Pure Appl. Math., 35 (1982),<br />

451–485.<br />

12 Emanuel, K. A.: <strong>Atmospheric</strong> Convection. Ox<strong>for</strong>d University Press,1994.<br />

13 Embid, P.; Majda, A. J.: Averaging over fast gravity waves <strong>for</strong> geophysical flows with arbitrary potential vorticity.<br />

Communications in Partial Differential Equations, 21 (1996),619–658.<br />

14 Embid, P.; Majda, A. J.: Averaging over fast gravity waves <strong>for</strong> geophysical flows with unbalanced initial data.<br />

Theoretical <strong>and</strong> Computational Fluid Dynamics, 11 (1998),155–169.<br />

15 Geratz, K. J.: Erweiterung eines Godunov-Typ-Verfahrens für mehrdimensionale kompressible Strömungen auf die<br />

Fälle kleiner und verschwindender Machzahl. PhD thesis,RWTH Aachen,1998.<br />

16 Geratz, K.; Klein, R.; Munz, C.-D.; Roller, S.: Multiple Pressure Variable (MPV) Approach <strong>for</strong> Low Mach Number<br />

<strong>Flows</strong> Based on <strong>Asymptotic</strong> Analysis. In: Hirschel, E. H. (ed.): Flow simulation with high-per<strong>for</strong>mance computers<br />

II. DFG priority research programme results. Notes on Numerical Fluid Mechanics,52. Vieweg Verlag,Braunschweig,<br />

1996.<br />

17 Gill, A. E. (ed.): Atmosphere-Ocean Dynamics. International Geophysical Series,30. Academic Press,London,1982.<br />

18 Guillard, H.; Viozat, C.: On the behoviour of upwind schemes in the low mach number limit. Computers <strong>and</strong> Fluids,<br />

28 (1999),63–86.<br />

19 Harlow, F. H.; Welch, J. E.: Numerical calculation of time dependent viscous incompressible flow. Phys. Fluids, 8<br />

(1965),2182–2189.<br />

20 Kaper, G.; Garbey, M. (eds.): <strong>Asymptotic</strong> analysis <strong>and</strong> the numerical solution of partial differential equations. Lecture<br />

Notes in Pure <strong>and</strong> Applied Mathematics,130. Marcel Dekker,New York,1991.<br />

21 Karki, K. C.; Patankar, S. V.: Pressure based calculation procedure <strong>for</strong> viscous flows at all speeds in arbitrary<br />

configurations. AIAA J., 27 (1989),1167–1174.<br />

22 Klainerman, S.; Majda, A. J.: Compressible <strong>and</strong> Incompressible fluids. Comm. Pure Appl. Math., 35 (1982),629–.<br />

23 Klein, R.: Semi-implicit extension of a godunov-type scheme based on low mach number asymptotics i: One-dimensional<br />

flow. Journal of Computational Physics, 121 (1995),213–237.<br />

24 Klein, R.; Botta, N.; Schneider, T.; Munz, C.; Roller, S.; Meister, A.; Hoffmann, L.; Sonar, T.: <strong>Asymptotic</strong><br />

adaptive methods <strong>for</strong> multi-scale problems in fluid mechanics. Journal of Engineering Mathematics, – (2000),—.


Klein, R.: Atmosphere <strong>Asymptotic</strong>s <strong>and</strong> Numerics – AUTHOR’S UPDATED PRINT 785<br />

submitted.<br />

25 Lai, M.; Bell, J. B.; Colella, P.: A projection method <strong>for</strong> combustion in the zero Mach number limit. AIAA Paper,<br />

3369 (1993).<br />

26 LeVeque, R. J.: A large time step generalization of Godunov’s method <strong>for</strong> systems of conservation laws. SIAM J.<br />

Num. Anal., 22 (1985),1051–1073.<br />

27 LeVeque, R. J.: Numerical Methods <strong>for</strong> Conservation Laws. Birkhäuser Verlag,Zürich,Schweiz,1992. IBSN 0-521-<br />

43009-7.<br />

28 Majda, A.; Sethian, J.: The derivation <strong>and</strong> numerical solution of the equations <strong>for</strong> zero mach number combustion.<br />

Combustion Science <strong>and</strong> Technology, 42 (1985),185–205.<br />

29 Merkle, C. L.; Choi, Y.-H.: Computation of low speed flow with heat addition. AIAA J., 25 (1987),831–838.<br />

30 Muraki, D. J.; Snyder, C.; Rotunno, R.: The next-order corrections to quasigeostrophic theory. Journal of the<br />

<strong>Atmospheric</strong> Sciences, 56 (1998),1547–1560.<br />

31 Ogura, Y.; Phillips, N. A.: Scale analysis of deep moist convection <strong>and</strong> some related numerical calculations. Journal<br />

of Atmosphere Science, 19 (1962),173–179.<br />

32 Pedlosky, J. (ed.): Geophysical Fluid Dynamics. Springer,2 edition,1987.<br />

33 Roller, S.; Munz, C.-D.; Geratz, K.; Klein, R.: The extension of incompressible flow solvers to the weakly<br />

compressible regime. Theoretical <strong>and</strong> Numerical Fluid Dynamics,(1999),submitted <strong>for</strong> publication.<br />

34 Schlichting, H. (ed.): Boundary Layer Theory. MacGraw-Hill,1968.<br />

35 Schneider, T.; Botta, N.; Klein, R.; Geratz, K. J.: Extension of finite volume compressible flow solvers to<br />

multi-dimensional,variable density zero mach number flows. Journal of Computational Physics,155 (1999),248–286.<br />

36 Schneider, W. (ed.): Mathematische Methoden in der Strömungsmechanik. Vieweg,1978.<br />

37 Schochet, S.: Fast singular limits of hyperbolic pdes. Journal of Differential Equations, 114 (1994),476–512.<br />

38 Sesterhenn, J.; Müller, B.; Thomann, H.: Computation of compressible low mach number flow. Computational<br />

Fluid Dynamics, 2 (1992),829–833.<br />

39 Smith, R. B.: Why can’t stably stratified air rise over high ground. In: Blumen, W. (ed.): <strong>Atmospheric</strong> processes<br />

over complex terrain. Meteorological Monographs,23,1990,pp. 105–107.<br />

40 Turkel, E.: Preconditioned Methods <strong>for</strong> Solving the Incompressible <strong>and</strong> Low Speed Compressible Equations. Journal<br />

of Computational Physics, 72 (1987),277 – 298.<br />

41 van Dyke, M. (ed.): Perturbation Methods in Fluid Mechanics,Annotated Ed. Parabolic Press,1975.<br />

42 Volpe, G.: On the use <strong>and</strong> accuracy of compressible flow codes at low mach numbers. AIAA Paper 91-1662,1991.<br />

43 Worlikar, A.; Knio, O.; Klein, R.: Numerical simulation of a thermo-acoustic refrigerator: Ii stratified flow around<br />

the stack. Journal of Computational Physics, 144 (1998),299–324.<br />

44 Zeytounian, R. K. (ed.): Meteorological Fluid Dynamics. Lecture Notes in Physics,m5. Springer,1991.<br />

Address: Prof. Rupert Klein,Potsdam Institut für Klimafolgen<strong>for</strong>schung,Telegrafenberg C4,14412 Potsdam,Germany,<br />

email: rupert.klein@pik-potsdam.de

Hooray! Your file is uploaded and ready to be published.

Saved successfully!

Ooh no, something went wrong!